Colloid

Last updated
SEM image of a colloid. SEM Image of Colloidal Particles.jpg
SEM image of a colloid.

A colloid is a mixture in which one substance consisting of microscopically dispersed insoluble particles is suspended throughout another substance. Some definitions specify that the particles must be dispersed in a liquid, [1] while others extend the definition to include substances like aerosols and gels. The term colloidal suspension refers unambiguously to the overall mixture (although a narrower sense of the word suspension is distinguished from colloids by larger particle size). A colloid has a dispersed phase (the suspended particles) and a continuous phase (the medium of suspension). The dispersed phase particles have a diameter of approximately 1 nanometre to 1 micrometre. [2] [3]

Contents

Some colloids are translucent because of the Tyndall effect, which is the scattering of light by particles in the colloid. Other colloids may be opaque or have a slight color.

Colloidal suspensions are the subject of interface and colloid science. This field of study began in 1845 by Francesco Selmi, [4] [5] [6] [7] who called them pseudosolutions, and expanded by Michael Faraday [8] and Thomas Graham, who coined the term colloid in 1861. [9]

IUPAC definition

Colloid: Short synonym for colloidal system. [10] [11]

Colloidal: State of subdivision such that the molecules or polymolecular particles dispersed in a medium have at least one dimension between approximately 1 nm and 1 μm, or that in a system discontinuities are found at distances of that order. [10] [11] [12]

Classification of colloids

Colloids can be classified as follows:

Medium/phaseDispersed phase
GasLiquidSolid
Dispersion
medium
GasNo such colloids are known.
Helium and xenon are known to be immiscible under certain conditions. [13] [14]
Liquid aerosol
Examples: fog, clouds, condensation, mist, steam, hair sprays
Solid aerosol
Examples: smoke, ice cloud, atmospheric particulate matter
Liquid Foam
Example: whipped cream, shaving cream
Emulsion or Liquid crystal
Examples: milk, mayonnaise, hand cream, latex, biological membranes , liquid biomolecular condensate
Sol
Examples: pigmented ink, sediment, precipitates, solid biomolecular condensate
SolidSolid foam
Examples: aerogel, floating soap, styrofoam, pumice
Gel
Examples: agar, gelatin, jelly, gel-like biomolecular condensate
Solid sol
Example: cranberry glass

Homogeneous mixtures with a dispersed phase in this size range may be called colloidal aerosols, colloidal emulsions, colloidal suspensions, colloidal foams, colloidal dispersions, or hydrosols.

Hydrocolloids

Hydrocolloids describe certain chemicals (mostly polysaccharides and proteins) that are colloidally dispersible in water. Thus becoming effectively "soluble" they change the rheology of water by raising the viscosity and/or inducing gelation. They may provide other interactive effects with other chemicals, in some cases synergistic, in others antagonistic. Using these attributes hydrocolloids are very useful chemicals since in many areas of technology from foods through pharmaceuticals, personal care and industrial applications, they can provide stabilization, destabilization and separation, gelation, flow control, crystallization control and numerous other effects. Apart from uses of the soluble forms some of the hydrocolloids have additional useful functionality in a dry form if after solubilization they have the water removed - as in the formation of films for breath strips or sausage casings or indeed, wound dressing fibers, some being more compatible with skin than others. There are many different types of hydrocolloids each with differences in structure function and utility that generally are best suited to particular application areas in the control of rheology and the physical modification of form and texture. Some hydrocolloids like starch and casein are useful foods as well as rheology modifiers, others have limited nutritive value, usually providing a source of fiber. [15]

The term hydrocolloids also refers to a type of dressing designed to lock moisture in the skin and help the natural healing process of skin to reduce scarring, itching and soreness.

Components

Hydrocolloids contain some type of gel-forming agent, such as sodium carboxymethylcellulose (NaCMC) and gelatin. They are normally combined with some type of sealant, i.e. polyurethane to 'stick' to the skin.

Colloid compared with solution

A colloid has a dispersed phase and a continuous phase, whereas in a solution, the solute and solvent constitute only one phase. A solute in a solution are individual molecules or ions, whereas colloidal particles are bigger. For example, in a solution of salt in water, the sodium chloride (NaCl) crystal dissolves, and the Na+ and Cl ions are surrounded by water molecules.  However, in a colloid such as milk, the colloidal particles are globules of fat, rather than individual fat molecules. Because colloid is multiple phases, it has very different properties compared to fully mixed, continuous solution. [16]

Interaction between particles

The following forces play an important role in the interaction of colloid particles: [17] [18]

Sedimentation velocity

Brownian motion of 350 nm diameter polymer colloidal particles. Brownian Motion.gif
Brownian motion of 350 nm diameter polymer colloidal particles.

The Earth’s gravitational field acts upon colloidal particles. Therefore, if the colloidal particles are denser than the medium of suspension, they will sediment (fall to the bottom), or if they are less dense, they will cream (float to the top). Larger particles also have a greater tendency to sediment because they have smaller Brownian motion to counteract this movement.

The sedimentation or creaming velocity is found by equating the Stokes drag force with the gravitational force:

where

is the Archimedean weight of the colloidal particles,
is the viscosity of the suspension medium,
is the radius of the colloidal particle,

and is the sedimentation or creaming velocity.

The mass of the colloidal particle is found using:

where

is the volume of the colloidal particle, calculated using the volume of a sphere ,

and is the difference in mass density between the colloidal particle and the suspension medium.

By rearranging, the sedimentation or creaming velocity is:

There is an upper size-limit for the diameter of colloidal particles because particles larger than 1 μm tend to sediment, and thus the substance would no longer be considered a colloidal suspension. [19]

The colloidal particles are said to be in sedimentation equilibrium if the rate of sedimentation is equal to the rate of movement from Brownian motion.

Preparation

There are two principal ways to prepare colloids: [20]

Stabilization

The stability of a colloidal system is defined by particles remaining suspended in solution and depends on the interaction forces between the particles. These include electrostatic interactions and van der Waals forces, because they both contribute to the overall free energy of the system. [21]

A colloid is stable if the interaction energy due to attractive forces between the colloidal particles is less than kT, where k is the Boltzmann constant and T is the absolute temperature. If this is the case, then the colloidal particles will repel or only weakly attract each other, and the substance will remain a suspension.

If the interaction energy is greater than kT, the attractive forces will prevail, and the colloidal particles will begin to clump together. This process is referred to generally as aggregation, but is also referred to as flocculation, coagulation or precipitation. [22] While these terms are often used interchangeably, for some definitions they have slightly different meanings. For example, coagulation can be used to describe irreversible, permanent aggregation where the forces holding the particles together are stronger than any external forces caused by stirring or mixing. Flocculation can be used to describe reversible aggregation involving weaker attractive forces, and the aggregate is usually called a floc. The term precipitation is normally reserved for describing a phase change from a colloid dispersion to a solid (precipitate) when it is subjected to a perturbation. [19] Aggregation causes sedimentation or creaming, therefore the colloid is unstable: if either of these processes occur the colloid will no longer be a suspension.

Examples of a stable and of an unstable colloidal dispersion. ColloidalStability.png
Examples of a stable and of an unstable colloidal dispersion.

Electrostatic stabilization and steric stabilization are the two main mechanisms for stabilization against aggregation.

A combination of the two mechanisms is also possible (electrosteric stabilization).

Steric and gel network stabilization. ComparisonStericStab-ShearThinningFluids2.png
Steric and gel network stabilization.

A method called gel network stabilization represents the principal way to produce colloids stable to both aggregation and sedimentation. The method consists in adding to the colloidal suspension a polymer able to form a gel network. Particle settling is hindered by the stiffness of the polymeric matrix where particles are trapped, [26] and the long polymeric chains can provide a steric or electrosteric stabilization to dispersed particles. Examples of such substances are xanthan and guar gum.

Destabilization

Destabilization can be accomplished by different methods:

Unstable colloidal suspensions of low-volume fraction form clustered liquid suspensions, wherein individual clusters of particles sediment if they are more dense than the suspension medium, or cream if they are less dense. However, colloidal suspensions of higher-volume fraction form colloidal gels with viscoelastic properties. Viscoelastic colloidal gels, such as bentonite and toothpaste, flow like liquids under shear, but maintain their shape when shear is removed. It is for this reason that toothpaste can be squeezed from a toothpaste tube, but stays on the toothbrush after it is applied.

Monitoring stability

Measurement principle of multiple light scattering coupled with vertical scanning MLS scan.gif
Measurement principle of multiple light scattering coupled with vertical scanning

The most widely used technique to monitor the dispersion state of a product, and to identify and quantify destabilization phenomena, is multiple light scattering coupled with vertical scanning. [28] [29] [30] [31] This method, known as turbidimetry, is based on measuring the fraction of light that, after being sent through the sample, it backscattered by the colloidal particles. The backscattering intensity is directly proportional to the average particle size and volume fraction of the dispersed phase. Therefore, local changes in concentration caused by sedimentation or creaming, and clumping together of particles caused by aggregation, are detected and monitored. [32] These phenomena are associated with unstable colloids.

Dynamic light scattering can be used to detect the size of a colloidal particle by measuring how fast they diffuse. This method involves directing laser light towards a colloid. The scattered light will form an interference pattern, and the fluctuation in light intensity in this pattern is caused by the Brownian motion of the particles. If the apparent size of the particles increases due to them clumping together via aggregation, it will result in slower Brownian motion. This technique can confirm that aggregation has occurred if the apparent particle size is determined to be beyond the typical size range for colloidal particles. [21]

Accelerating methods for shelf life prediction

The kinetic process of destabilisation can be rather long (up to several months or years for some products). Thus, it is often required for the formulator to use further accelerating methods to reach reasonable development time for new product design. Thermal methods are the most commonly used and consist of increasing temperature to accelerate destabilisation (below critical temperatures of phase inversion or chemical degradation). Temperature affects not only viscosity, but also interfacial tension in the case of non-ionic surfactants or more generally interactions forces inside the system. Storing a dispersion at high temperatures enables to simulate real life conditions for a product (e.g. tube of sunscreen cream in a car in the summer), but also to accelerate destabilisation processes up to 200 times. Mechanical acceleration including vibration, centrifugation and agitation are sometimes used. They subject the product to different forces that pushes the particles / droplets against one another, hence helping in the film drainage. Some emulsions would never coalesce in normal gravity, while they do under artificial gravity. [33] Segregation of different populations of particles have been highlighted when using centrifugation and vibration. [34]

As a model system for atoms

In physics, colloids are an interesting model system for atoms. [35] Micrometre-scale colloidal particles are large enough to be observed by optical techniques such as confocal microscopy. Many of the forces that govern the structure and behavior of matter, such as excluded volume interactions or electrostatic forces, govern the structure and behavior of colloidal suspensions. For example, the same techniques used to model ideal gases can be applied to model the behavior of a hard sphere colloidal suspension. Phase transitions in colloidal suspensions can be studied in real time using optical techniques, [36] and are analogous to phase transitions in liquids. In many interesting cases optical fluidity is used to control colloid suspensions. [36] [37]

Crystals

A colloidal crystal is a highly ordered array of particles that can be formed over a very long range (typically on the order of a few millimeters to one centimeter) and that appear analogous to their atomic or molecular counterparts. [38] One of the finest natural examples of this ordering phenomenon can be found in precious opal, in which brilliant regions of pure spectral color result from close-packed domains of amorphous colloidal spheres of silicon dioxide (or silica, SiO2). [39] [40] These spherical particles precipitate in highly siliceous pools in Australia and elsewhere, and form these highly ordered arrays after years of sedimentation and compression under hydrostatic and gravitational forces. The periodic arrays of submicrometre spherical particles provide similar arrays of interstitial voids, which act as a natural diffraction grating for visible light waves, particularly when the interstitial spacing is of the same order of magnitude as the incident lightwave. [41] [42]

Thus, it has been known for many years that, due to repulsive Coulombic interactions, electrically charged macromolecules in an aqueous environment can exhibit long-range crystal-like correlations with interparticle separation distances, often being considerably greater than the individual particle diameter. In all of these cases in nature, the same brilliant iridescence (or play of colors) can be attributed to the diffraction and constructive interference of visible lightwaves that satisfy Bragg’s law, in a matter analogous to the scattering of X-rays in crystalline solids.

The large number of experiments exploring the physics and chemistry of these so-called "colloidal crystals" has emerged as a result of the relatively simple methods that have evolved in the last 20 years for preparing synthetic monodisperse colloids (both polymer and mineral) and, through various mechanisms, implementing and preserving their long-range order formation. [43]

In biology

Colloidal phase separation is an important organising principle for compartmentalisation of both the cytoplasm and nucleus of cells into biomolecular condensates —similar in importance to compartmentalisation via lipid bilayer membranes, a type of liquid crystal. The term biomolecular condensate has been used to refer to clusters of macromolecules that arise via liquid-liquid or liquid-solid phase separation within cells. Macromolecular crowding strongly enhances colloidal phase separation and formation of biomolecular condensates.

In the environment

Colloidal particles can also serve as transport vector [44] of diverse contaminants in the surface water (sea water, lakes, rivers, fresh water bodies) and in underground water circulating in fissured rocks [45] (e.g. limestone, sandstone, granite). Radionuclides and heavy metals easily sorb onto colloids suspended in water. Various types of colloids are recognised: inorganic colloids (e.g. clay particles, silicates, iron oxy-hydroxides), organic colloids (humic and fulvic substances). When heavy metals or radionuclides form their own pure colloids, the term " eigencolloid " is used to designate pure phases, i.e., pure Tc(OH)4, U(OH)4, or Am(OH)3. Colloids have been suspected for the long-range transport of plutonium on the Nevada Nuclear Test Site. They have been the subject of detailed studies for many years. However, the mobility of inorganic colloids is very low in compacted bentonites and in deep clay formations [46] because of the process of ultrafiltration occurring in dense clay membrane. [47] The question is less clear for small organic colloids often mixed in porewater with truly dissolved organic molecules. [48]

In soil science, the colloidal fraction in soils consists of tiny clay and humus particles that are less than 1μm in diameter and carry either positive and/or negative electrostatic charges that vary depending on the chemical conditions of the soil sample, i.e. soil pH. [49]

Intravenous therapy

Colloid solutions used in intravenous therapy belong to a major group of volume expanders, and can be used for intravenous fluid replacement. Colloids preserve a high colloid osmotic pressure in the blood, [50] and therefore, they should theoretically preferentially increase the intravascular volume, whereas other types of volume expanders called crystalloids also increase the interstitial volume and intracellular volume. However, there is still controversy to the actual difference in efficacy by this difference, [50] and much of the research related to this use of colloids is based on fraudulent research by Joachim Boldt. [51] Another difference is that crystalloids generally are much cheaper than colloids. [50]

Related Research Articles

<span class="mw-page-title-main">Emulsion</span> Mixture of two or more immiscible liquids

An emulsion is a mixture of two or more liquids that are normally immiscible owing to liquid-liquid phase separation. Emulsions are part of a more general class of two-phase systems of matter called colloids. Although the terms colloid and emulsion are sometimes used interchangeably, emulsion should be used when both phases, dispersed and continuous, are liquids. In an emulsion, one liquid is dispersed in the other. Examples of emulsions include vinaigrettes, homogenized milk, liquid biomolecular condensates, and some cutting fluids for metal working.

In polymer chemistry, emulsion polymerization is a type of radical polymerization that usually starts with an emulsion incorporating water, monomers, and surfactants. The most common type of emulsion polymerization is an oil-in-water emulsion, in which droplets of monomer are emulsified in a continuous phase of water. Water-soluble polymers, such as certain polyvinyl alcohols or hydroxyethyl celluloses, can also be used to act as emulsifiers/stabilizers. The name "emulsion polymerization" is a misnomer that arises from a historical misconception. Rather than occurring in emulsion droplets, polymerization takes place in the latex/colloid particles that form spontaneously in the first few minutes of the process. These latex particles are typically 100 nm in size, and are made of many individual polymer chains. The particles are prevented from coagulating with each other because each particle is surrounded by the surfactant ('soap'); the charge on the surfactant repels other particles electrostatically. When water-soluble polymers are used as stabilizers instead of soap, the repulsion between particles arises because these water-soluble polymers form a 'hairy layer' around a particle that repels other particles, because pushing particles together would involve compressing these chains.

<span class="mw-page-title-main">Suspension (chemistry)</span> Heterogeneous mixture of solid particles dispersed in a medium

In chemistry, a suspension is a heterogeneous mixture of a fluid that contains solid particles sufficiently large for sedimentation. The particles may be visible to the naked eye, usually must be larger than one micrometer, and will eventually settle, although the mixture is only classified as a suspension when and while the particles have not settled out.

<span class="mw-page-title-main">Dilatant</span> Material in which viscosity increases with the rate of shear strain

A dilatant material is one in which viscosity increases with the rate of shear strain. Such a shear thickening fluid, also known by the initialism STF, is an example of a non-Newtonian fluid. This behaviour is usually not observed in pure materials, but can occur in suspensions.

The DLVO theory explains the aggregation and kinetic stability of aqueous dispersions quantitatively and describes the force between charged surfaces interacting through a liquid medium. It combines the effects of the van der Waals attraction and the electrostatic repulsion due to the so-called double layer of counterions. The electrostatic part of the DLVO interaction is computed in the mean field approximation in the limit of low surface potentials - that is when the potential energy of an elementary charge on the surface is much smaller than the thermal energy scale, . For two spheres of radius each having a charge separated by a center-to-center distance in a fluid of dielectric constant containing a concentration of monovalent ions, the electrostatic potential takes the form of a screened-Coulomb or Yukawa potential,

<span class="mw-page-title-main">Polyelectrolyte</span> Polymers whose repeating units bear an electrolyte group

Polyelectrolytes are polymers whose repeating units bear an electrolyte group. Polycations and polyanions are polyelectrolytes. These groups dissociate in aqueous solutions (water), making the polymers charged. Polyelectrolyte properties are thus similar to both electrolytes (salts) and polymers and are sometimes called polysalts. Like salts, their solutions are electrically conductive. Like polymers, their solutions are often viscous. Charged molecular chains, commonly present in soft matter systems, play a fundamental role in determining structure, stability and the interactions of various molecular assemblies. Theoretical approaches to describe their statistical properties differ profoundly from those of their electrically neutral counterparts, while technological and industrial fields exploit their unique properties. Many biological molecules are polyelectrolytes. For instance, polypeptides, glycosaminoglycans, and DNA are polyelectrolytes. Both natural and synthetic polyelectrolytes are used in a variety of industries.

<span class="mw-page-title-main">Flocculation</span> Process by which colloidal particles come out of suspension to precipitate as floc or flake

In colloidal chemistry, flocculation is a process by which colloidal particles come out of suspension to sediment in the form of floc or flake, either spontaneously or due to the addition of a clarifying agent. The action differs from precipitation in that, prior to flocculation, colloids are merely suspended, under the form of a stable dispersion and are not truly dissolved in solution.

In materials science, the sol–gel process is a method for producing solid materials from small molecules. The method is used for the fabrication of metal oxides, especially the oxides of silicon (Si) and titanium (Ti). The process involves conversion of monomers into a colloidal solution (sol) that acts as the precursor for an integrated network of either discrete particles or network polymers. Typical precursors are metal alkoxides. Sol–gel process is used to produce ceramic nanoparticles.

A surface charge is an electric charge present on a two-dimensional surface. These electric charges are constrained on this 2-D surface, and surface charge density, measured in coulombs per square meter (C•m−2), is used to describe the charge distribution on the surface. The electric potential is continuous across a surface charge and the electric field is discontinuous, but not infinite; this is unless the surface charge consists of a dipole layer. In comparison, the potential and electric field both diverge at any point charge or linear charge.

<span class="mw-page-title-main">Particle aggregation</span> Clumping of particles in suspension

Particle agglomeration refers to the formation of assemblages in a suspension and represents a mechanism leading to the functional destabilization of colloidal systems. During this process, particles dispersed in the liquid phase stick to each other, and spontaneously form irregular particle assemblages, flocs, or agglomerates. This phenomenon is also referred to as coagulation or flocculation and such a suspension is also called unstable. Particle agglomeration can be induced by adding salts or other chemicals referred to as coagulant or flocculant.

A dispersion is a system in which distributed particles of one material are dispersed in a continuous phase of another material. The two phases may be in the same or different states of matter.

<span class="mw-page-title-main">Suspension polymerization</span> Polymerization reaction among monomers suspended in a liquid

In polymer chemistry, suspension polymerization is a heterogeneous radical polymerization process that uses mechanical agitation to mix a monomer or mixture of monomers in a liquid phase, such as water, while the monomers polymerize, forming spheres of polymer. The monomer droplets are suspended in the liquid phase. The individual monomer droplets can be considered as undergoing bulk polymerization. The liquid phase outside these droplets help in better conduction of heat and thus tempering the increase in temperature.

Sedimentation potential occurs when dispersed particles move under the influence of either gravity or centrifugation or electricity in a medium. This motion disrupts the equilibrium symmetry of the particle's double layer. While the particle moves, the ions in the electric double layer lag behind due to the liquid flow. This causes a slight displacement between the surface charge and the electric charge of the diffuse layer. As a result, the moving particle creates a dipole moment. The sum of all of the dipoles generates an electric field which is called sedimentation potential. It can be measured with an open electrical circuit, which is also called sedimentation current.

<span class="mw-page-title-main">Colloidal crystal</span> Ordered array of colloidal particles

A colloidal crystal is an ordered array of colloidal particles and fine grained materials analogous to a standard crystal whose repeating subunits are atoms or molecules. A natural example of this phenomenon can be found in the gem opal, where spheres of silica assume a close-packed locally periodic structure under moderate compression. Bulk properties of a colloidal crystal depend on composition, particle size, packing arrangement, and degree of regularity. Applications include photonics, materials processing, and the study of self-assembly and phase transitions.

Colloidal silicas are suspensions of fine amorphous, nonporous, and typically spherical silica particles in a liquid phase. It may be produced by Stöber process from Tetraethyl orthosilicate (TEOS).

Adsorption of polyelectrolytes on solid substrates is a surface phenomenon where long-chained polymer molecules with charged groups bind to a surface that is charged in the opposite polarity. On the molecular level, the polymers do not actually bond to the surface, but tend to "stick" to the surface via intermolecular forces and the charges created by the dissociation of various side groups of the polymer. Because the polymer molecules are so long, they have a large amount of surface area with which to contact the surface and thus do not desorb as small molecules are likely to do. This means that adsorbed layers of polyelectrolytes form a very durable coating. Due to this important characteristic of polyelectrolyte layers they are used extensively in industry as flocculants, for solubilization, as supersorbers, antistatic agents, as oil recovery aids, as gelling aids in nutrition, additives in concrete, or for blood compatibility enhancement to name a few.

Clarifying agents are used to remove suspended solids from liquids by inducing flocculation, causing the solids to form larger aggregates that can be easily removed after they either float to the surface or sink to the bottom of the containment vessel.

Macroemulsions are dispersed liquid-liquid, thermodynamically unstable systems with particle sizes ranging from 1 to 100 μm, which, most often, do not form spontaneously. Macroemulsions scatter light effectively and therefore appear milky, because their droplets are greater than a wavelength of light. They are part of a larger family of emulsions along with miniemulsions. As with all emulsions, one phase serves as the dispersing agent. It is often called the continuous or outer phase. The remaining phase(s) are disperse or inner phase(s), because the liquid droplets are finely distributed amongst the larger continuous phase droplets. This type of emulsion is thermodynamically unstable, but can be stabilized for a period of time with applications of kinetic energy. Surfactants are used to reduce the interfacial tension between the two phases, and induce macroemulsion stability for a useful amount of time. Emulsions can be stabilized otherwise with polymers, solid particles or proteins.

Polyelectrolytes are charged polymers capable of stabilizing colloidal emulsions through electrostatic interactions. Their effectiveness can be dependent on molecular weight, pH, solvent polarity, ionic strength, and the hydrophilic-lipophilic balance (HLB). Stabilized emulsions are useful in many industrial processes, including deflocculation, drug delivery, petroleum waste treatment, and food technology.

A depletion force is an effective attractive force that arises between large colloidal particles that are suspended in a dilute solution of depletants, which are smaller solutes that are preferentially excluded from the vicinity of the large particles. One of the earliest reports of depletion forces that lead to particle coagulation is that of Bondy, who observed the separation or "creaming" of rubber latex upon addition of polymer depletant molecules to solution. More generally, depletants can include polymers, micelles, osmolytes, ink, mud, or paint dispersed in a continuous phase.

References

  1. 1 2 Israelachvili, Jacob N. (2011). Intermolecular and surface forces (4rd ed.). Burlington, MA: Academic Press. ISBN   978-0-08-092363-5. OCLC   706803091.
  2. International Union of Pure and Applied Chemistry. Subcommittee on Polymer Terminology; Jones, Richard G. (2009). Compendium of polymer terminology and nomenclature : IUPAC recommendations, 2008. Cambridge: Royal Society of Chemistry. ISBN   978-1-84755-942-5. OCLC   406528399.
  3. Stepto, Robert F. T. (1 January 2009). "Dispersity in polymer science (IUPAC Recommendations 2009)". Pure and Applied Chemistry. 81 (2): 351–353. doi: 10.1351/PAC-REC-08-05-02 . S2CID   95122531.
  4. Selmi, Francesco "Studi sulla dimulsione di cloruro d'argento". Nuovi Annali delle Scienze Naturali di Bologna, 1845.
  5. Selmi, Francesco, Studio intorno alle pseudo-soluzioni degli azzurri di Prussia ed alla influenza dei sali nel guastarle, Bologna: Tipi Sassi, 1847
  6. Hatschek, Emil, The Foundations of Colloid Chemistry, A selection of early papers bearing on the subject, The British Association Committee on Colloid Chemistry, London, 1925
  7. Selmi, Francesco - Sur le soufre pseudosoluble, sa pseudosolution e le soufre mou, Journal de Pharmacie et de Chimie, tome 21, 1852, Paris
  8. Tweney, Ryan D. (2006). "Discovering Discovery: How Faraday Found the First Metallic Colloid". Perspectives on Science. 14: 97–121. doi:10.1162/posc.2006.14.1.97. S2CID   55882753.
  9. "X. Liquid diffusion applied to analysis". Philosophical Transactions of the Royal Society of London. 151: 183–224. 1861. doi:10.1098/rstl.1861.0011. S2CID   186208563.. Page 183: "As gelatine appears to be its type, it is proposed to designate substances of the class as colloids, and to speak of their peculiar form of aggregation as the colloidal condition of matter."
  10. 1 2 Richard G. Jones; Edward S. Wilks; W. Val Metanomski; Jaroslav Kahovec; Michael Hess; Robert Stepto; Tatsuki Kitayama, eds. (2009). Compendium of Polymer Terminology and Nomenclature (IUPAC Recommendations 2008) (2nd ed.). RSC Publ. p. 464. ISBN   978-0-85404-491-7.
  11. 1 2 Stepto, Robert F. T. (2009). "Dispersity in polymer science (IUPAC Recommendations 2009)" (PDF). Pure and Applied Chemistry . 81 (2): 351–353. doi:10.1351/PAC-REC-08-05-02. S2CID   95122531. Archived (PDF) from the original on 9 October 2022.
  12. Slomkowski, Stanislaw; Alemán, José V.; Gilbert, Robert G.; Hess, Michael; Horie, Kazuyuki; Jones, Richard G.; Kubisa, Przemyslaw; Meisel, Ingrid; Mormann, Werner; Penczek, Stanisław; Stepto, Robert F. T. (2011). "Terminology of polymers
    and polymerization processes in dispersed systems (IUPAC Recommendations 2011)"
    (PDF). Pure and Applied Chemistry . 83 (12): 2229–2259. doi:10.1351/PAC-REC-10-06-03. S2CID   96812603. Archived (PDF) from the original on 9 October 2022.
  13. de Swaan Arons, J.; Diepen, G. A. M. (2010). "Immiscibility of gases. The system He-Xe: (Short communication)". Recueil des Travaux Chimiques des Pays-Bas. 82 (8): 806. doi:10.1002/recl.19630820810.
  14. de Swaan Arons, J.; Diepen, G. A. M. (1966). "Gas—Gas Equilibria". J. Chem. Phys. 44 (6): 2322. Bibcode:1966JChPh..44.2322D. doi:10.1063/1.1727043.
  15. Saha, Dipjyoti; Bhattacharya, Suvendu (6 November 2010). "Hydrocolloids as thickening and gelling agents in food: a critical review". Journal of Food Science and Technology . 47 (6): 587–597. doi:10.1007/s13197-010-0162-6. PMC   3551143 . PMID   23572691.
  16. McBride, Samantha A.; Skye, Rachael; Varanasi, Kripa K. (2020). "Differences between Colloidal and Crystalline Evaporative Deposits". Langmuir. 36 (40): 11732–11741. doi:10.1021/acs.langmuir.0c01139. PMID   32937070. S2CID   221770585.
  17. Lekkerkerker, Henk N.W.; Tuinier, Remco (2011). Colloids and the Depletion Interaction. Heidelberg: Springer. doi:10.1007/978-94-007-1223-2. ISBN   9789400712225. Archived from the original on 14 April 2019. Retrieved 5 September 2018.
  18. van Anders, Greg; Klotsa, Daphne; Ahmed, N. Khalid; Engel, Michael; Glotzer, Sharon C. (2014). "Understanding shape entropy through local dense packing". Proc Natl Acad Sci USA. 111 (45): E4812–E4821. arXiv: 1309.1187 . Bibcode:2014PNAS..111E4812V. doi: 10.1073/pnas.1418159111 . PMC   4234574 . PMID   25344532.
  19. 1 2 3 Cosgrove, Terence (2010). Colloid Science: Principles, Methods and Applications. John Wiley & Sons. ISBN   9781444320183.
  20. Kopeliovich, Dmitri. Preparation of colloids. substech.com
  21. 1 2 Everett, D. H. (1988). Basic principles of colloid science. London: Royal Society of Chemistry. ISBN   978-1-84755-020-0. OCLC   232632488.
  22. Slomkowski, Stanislaw; Alemán, José V.; Gilbert, Robert G.; Hess, Michael; Horie, Kazuyuki; Jones, Richard G.; Kubisa, Przemyslaw; Meisel, Ingrid; Mormann, Werner; Penczek, Stanisław; Stepto, Robert F. T. (10 September 2011). "Terminology of polymers and polymerization processes in dispersed systems (IUPAC Recommendations 2011)". Pure and Applied Chemistry (in German). 83 (12): 2229–2259. doi: 10.1351/PAC-REC-10-06-03 . S2CID   96812603.
  23. Park, Soo-Jin; Seo, Min-Kang (1 January 2011). "Intermolecular Force". Interface Science and Technology. 18: 1–57. doi:10.1016/B978-0-12-375049-5.00001-3. ISBN   9780123750495.
  24. Tadros, Tharwat F. (2007). Colloid stability : the role of surface forces. Part I. Weinheim: Wiley-VCH. ISBN   978-3-527-63107-0. OCLC   701308697.
  25. Genz, Ulrike; D'Aguanno, Bruno; Mewis, Jan; Klein, Rudolf (1 July 1994). "Structure of Sterically Stabilized Colloids". Langmuir. 10 (7): 2206–2212. doi:10.1021/la00019a029.
  26. Comba, Silvia; Sethi (August 2009). "Stabilization of highly concentrated suspensions of iron nanoparticles using shear-thinning gels of xanthan gum". Water Research. 43 (15): 3717–3726. Bibcode:2009WatRe..43.3717C. doi:10.1016/j.watres.2009.05.046. PMID   19577785.
  27. Bean, Elwood L.; Campbell, Sylvester J.; Anspach, Frederick R.; Ockershausen, Richard W.; Peterman, Charles J. (1964). "Zeta Potential Measurements in the Control of Coagulation Chemical Doses [with Discussion]". Journal (American Water Works Association). 56 (2): 214–227. doi:10.1002/j.1551-8833.1964.tb01202.x. JSTOR   41264141.
  28. Roland, I; Piel, G; Delattre, L; Evrard, B (2003). "Systematic characterisation of oil-in-water emulsions for formulation design". International Journal of Pharmaceutics. 263 (1–2): 85–94. doi:10.1016/S0378-5173(03)00364-8. PMID   12954183.
  29. Lemarchand, Caroline; Couvreur, Patrick; Besnard, Madeleine; Costantini, Dominique; Gref, Ruxandra (2003). "Novel polyester-polysaccharide nanoparticles". Pharmaceutical Research. 20 (8): 1284–92. doi:10.1023/A:1025017502379. PMID   12948027. S2CID   24157992.
  30. Mengual, O (1999). "Characterisation of instability of concentrated dispersions by a new optical analyser: the TURBISCAN MA 1000". Colloids and Surfaces A: Physicochemical and Engineering Aspects. 152 (1–2): 111–123. doi:10.1016/S0927-7757(98)00680-3.
  31. Bru, P.; et al. (2004). T. Provder; J. Texter (eds.). Particle sizing and characterisation.
  32. Matusiak, Jakub; Grządka, Elżbieta (8 December 2017). "Stability of colloidal systems - a review of the stability measurements methods". Annales Universitatis Mariae Curie-Sklodowska, sectio AA – Chemia. 72 (1): 33. doi: 10.17951/aa.2017.72.1.33 .
  33. Salager, J-L (2000). Françoise Nielloud; Gilberte Marti-Mestres (eds.). Pharmaceutical emulsions and suspensions. CRC press. p. 89. ISBN   978-0-8247-0304-2.
  34. Snabre, Patrick; Pouligny, Bernard (2008). "Size Segregation in a Fluid-like or Gel-like Suspension Settling under Gravity or in a Centrifuge". Langmuir. 24 (23): 13338–47. doi:10.1021/la802459u. PMID   18986182.
  35. Manoharan, Vinothan N. (2015). "Colloidal matter: Packing, geometry, and entropy" (PDF). Science. 349 (6251): 1253751. doi: 10.1126/science.1253751 . PMID   26315444. S2CID   5727282.
  36. 1 2 Greenfield, Elad; Nemirovsky, Jonathan; El-Ganainy, Ramy; Christodoulides, Demetri N; Segev, Mordechai (2013). "Shockwave based nonlinear optical manipulation in densely scattering opaque suspensions". Optics Express. 21 (20): 23785–23802. Bibcode:2013OExpr..2123785G. doi: 10.1364/OE.21.023785 . PMID   24104290.
  37. Greenfield, Elad; Rotschild, Carmel; Szameit, Alexander; Nemirovsky, Jonathan; El-Ganainy, Ramy; Christodoulides, Demetrios N; Saraf, Meirav; Lifshitz, Efrat; Segev, Mordechai (2011). "Light-induced self-synchronizing flow patterns". New Journal of Physics. 13 (5): 053021. Bibcode:2011NJPh...13e3021G. doi: 10.1088/1367-2630/13/5/053021 .
  38. Pieranski, P. (1983). "Colloidal Crystals". Contemporary Physics. 24: 25–73. Bibcode:1983ConPh..24...25P. doi:10.1080/00107518308227471.
  39. Sanders, J.V.; Sanders, J. V.; Segnit, E. R. (1964). "Structure of Opal". Nature. 204 (4962): 1151. Bibcode:1964Natur.204..990J. doi:10.1038/204990a0. S2CID   4191566.
  40. Darragh, P.J.; et al. (1976). "Opals". Scientific American. 234 (4): 84–95. Bibcode:1976SciAm.234d..84D. doi:10.1038/scientificamerican0476-84.
  41. Luck, Werner; Klier, Manfred; Wesslau, Hermann (1963). "Über Bragg-Reflexe mit sichtbarem Licht an monodispersen Kunststofflatices. II". Berichte der Bunsengesellschaft für Physikalische Chemie. 67 (1): 84–85. doi:10.1002/bbpc.19630670114.
  42. Hiltner, P.A.; Krieger, I.M. (1969). "Diffraction of light by ordered suspensions". J. Phys. Chem. 73 (7): 2306. doi:10.1021/j100727a049.
  43. Liu, Xuesong; Li, Zejing; Tang, Jianguo; Yu, Bing; Cong, Hailin (9 September 2013). "Current status and future developments in preparation and application of colloidal crystals". Chemical Society Reviews. 42 (19): 7774–7800. doi:10.1039/C3CS60078E. PMID   23836297.
  44. Frimmel, Fritz H.; Frank von der Kammer; Hans-Curt Flemming (2007). Colloidal transport in porous media (1 ed.). Springer. p. 292. ISBN   978-3-540-71338-8.
  45. Alonso, U.; T. Missana; A. Patelli; V. Rigato (2007). "Bentonite colloid diffusion through the host rock of a deep geological repository". Physics and Chemistry of the Earth, Parts A/B/C. 32 (1–7): 469–476. Bibcode:2007PCE....32..469A. doi:10.1016/j.pce.2006.04.021.
  46. Voegelin, A.; Kretzschmar, R. (December 2002). "Stability and mobility of colloids in Opalinus Clay" (PDF). Technischer Bericht / NTB. Nagra Technical Report 02-14. Institute of Terrestrial Ecology, ETH Zürich: 47. ISSN   1015-2636. Archived from the original (PDF) on 9 March 2009. Retrieved 22 February 2009.
  47. "Diffusion of colloids in compacted bentonite". Archived from the original on 4 March 2009. Retrieved 12 February 2009.
  48. Wold, Susanna; Trygve Eriksen (2007). "Diffusion of humic colloids in compacted bentonite". Physics and Chemistry of the Earth, Parts A/B/C. 32 (1–7): 477–484. Bibcode:2007PCE....32..477W. doi:10.1016/j.pce.2006.05.002.
  49. Weil, Ray; Brady, Nyle C. (11 October 2018). Elements of the nature and properties of soils (Fourth ed.). New York, NY. ISBN   9780133254594. OCLC   1035317420.{{cite book}}: CS1 maint: location missing publisher (link)
  50. 1 2 3 Martin, Gregory S. (19 April 2005). "An Update on Intravenous Fluids". Medscape . Retrieved 6 July 2016.
  51. Blake, Heidi (3 March 2011). "Millions of surgery patients at risk in drug research fraud scandal". The Telegraph. UK. Archived from the original on 4 November 2011. Retrieved 4 November 2011.